Tuning the Wettability of Halloysite Clay Nanotubes by Surface

Dec 3, 2015 - The sign in the bracket (equation) is negative for particle desorption into water (0 ≤ θ ≤ 90°) and positive for particle desorpti...
5 downloads 0 Views 10MB Size
Article pubs.acs.org/Langmuir

Tuning the Wettability of Halloysite Clay Nanotubes by Surface Carbonization for Optimal Emulsion Stabilization Olasehinde Owoseni,† Yueheng Zhang,† Yang Su,† Jibao He,‡ Gary L. McPherson,§ Arijit Bose,⊥ and Vijay T. John*,†,∥ †

Department of Chemical and Biomolecular Engineering, ‡Coordinated Instrumentation Facility, §Department of Chemistry, and Vector Borne Infectious Disease Center, Tulane University, New Orleans, Louisiana 70118, United States ⊥ Department of Chemical Engineering, University of Rhode Island, Kingston, Rhode Island 02881, United States ∥

ABSTRACT: The carbonization of hydrophilic particle surfaces provides an effective route for tuning particle wettability in the preparation of particlestabilized emulsions. The wettability of naturally occurring halloysite clay nanotubes (HNT) is successfully tuned by the selective carbonization of the negatively charged external HNT surface. The positively charge chitosan biopolymer binds to the negatively charged external HNT surface by electrostatic attraction and hydrogen bonding, yielding carbonized halloysite nanotubes (CHNT) on pyrolysis in an inert atmosphere. Relative to the native HNT, the oil emulsification ability of the CHNT at intermediate levels of carbonization is significantly enhanced due to the thermodynamically more favorable attachment of the particles at the oil−water interface. Cryogenic scanning electron microscopy (cryo-SEM) imaging reveals that networks of CHNT attach to the oil−water interface with the particles in a side-on orientation. The concepts advanced here can be extended to other inorganic solids and carbon sources for the optimal design of particle-stabilized emulsions.

1. INTRODUCTION

Similar free energy analysis on a cylindrical particle at a planar oil−water interface gives the free energy change (ΔGw) on removing the particle from the interface into the water as4,8

Emulsions stabilized by fine solid particles are relevant in a wide range of processes such as in the food, pharmaceutical, and petroleum industries.1 Solid stabilized emulsions have also attracted interest in environmental remediation applications including the treatment of crude oil spills.2−5 A key characteristic of an effective emulsifier, including particles used as oil spill dispersants, is the ability to adsorb at the oil− water interface and inhibit droplet coalescence and phase separation.6−8 The particle wettability can be quantitatively expressed in terms of the three-phase contact angle which the particle makes at the oil−water interface.9 The attachment of particles to the oil−water interface with energies that are several orders of magnitude higher than the thermal energy provides a large steric hindrance to droplet coalescence and is one of the key underlying concepts behind particle stabilization of emulsions.10 Analysis on a spherical particle at a planar oil−water interface gives the work required to desorb the particle from the interface (ΔG) as2,9 ΔG = πrs 2γow(1 ± cos θ )2

⎡ ⎛ r⎞ ΔGw = 2rLγow ⎢sin θ − θ cos θ ⎜1 + ⎟ ⎝ L⎠ ⎣ +

for 0 ≤ θ ≤ 90° (2)

and the free energy change to remove the particle from the interface into the oil phase as ⎡ ⎛ r⎞ ΔG0 = 2rLγow ⎢sin θ + (π − θ ) cos θ ⎜1 + ⎟ ⎝ L⎠ ⎣ +

r cos 2 θ sin θ ⎤ ⎥ L ⎦

for 90° ≤ θ ≤ 180° (3)

where γow is the oil−water interfacial tension, r is the radius of the cylindrical particle, L is the particle length, and θ is the contact angle. Figure 1 shows the free energy of detachment from the oil−water interface for spherical and cylindrical particles of the same volume as a function of contact angle.

(1)

where γow is the oil−water interfacial tension, rs is the particle radius, and θ is the equilibrium contact angle measured through the water phase. The sign in the bracket (equation) is negative for particle desorption into water (0 ≤ θ ≤ 90°) and positive for particle desorption into oil (90° ≤ θ ≤ 180°).1 © 2015 American Chemical Society

r cos 2 θ sin θ ⎤ ⎥ L ⎦

Received: October 19, 2015 Revised: November 30, 2015 Published: December 3, 2015 13700

DOI: 10.1021/acs.langmuir.5b03878 Langmuir 2015, 31, 13700−13707

Article

Langmuir

surface of the silica particles was proposed to make the particles more hydrophobic and enhance their interfacial activity for oil emulsification.24 Recently we have advanced the use of naturally occurring halloysite aluminosilicate clay nanotube (HNT) for oil emulsification.4 An interesting aspect of the HNT is that it has a predominantly negatively charged outer silica surface and a positively charged inner alumina surface.4,25 The nanotubular morphology of the HNT has been exploited for the loading and delivery of surfactant to the oil−water interface.4,8,26 The effectiveness of particles in emulsion stabilization and interfacial delivery of materials are largely dependent on the partitioning of the HNT to the oil−water interface. Here we propose a new concept to optimize the attachment of the particles to the oil− water interface and tune the wettability of the hydrophilic native HNT by the systematic carbonization of the external surface of the particles. We exploit the selective binding of the readily available cationic biopolymer, chitosan, onto the negatively charged external surface of HNT27 to carbonize the HNT on pyrolysis in an inert atmosphere. We demonstrate that this is an effective route to tune the wettability of the nanotubes for optimal emulsion stabilization.

Figure 1. Free energy of particle detachment from the oil−water interface as a function of contact angle for a spherical particle (red curve) and cylindrical particles with aspect ratio of 5.0 and 10.0 (blue and green curves, respectively). All particles have the same volume of 7.85 × 10−21 m3, the oil−water interfacial tension was 49 mN/m, and the cylinder aspect ratio is defined as the length-to-diameter ratio.

A first observation is that the energy of attachment of the cylindrical particle at the oil−water interface is higher than for a spherical particle of the same volume, underscoring the role of particle shape on emulsion stabilization by the particles.1,11 For instance at θ = 90°, the work ΔG in eqs 1 and 2 reduces to the product of the interfacial tension (γow) and the area of the oil− water interface occupied by the particle corresponding to the area of the midsection of the particles. For a cylindrical particle with 50.0 nm radius and length of 1.0 μm, a spherical particle of equal volume will have a radius of 123.3 nm. The area of the midsection is calculated to be 1.0 × 10−13 and 4.8 × 10−14 m2 for the cylindrical and spherical particles, respectively, yielding a higher ΔG/kT value of 1.3 × 106 for the cylindrical particle compared to 6.2 × 105 for the spherical particle. The energy of attachment at the interface is higher for cylindrical particles with aspect ratio of 10 relative to cylindrical particles with an aspect ratio of 5. Madivala et al. experimentally characterized the role of particle shape and aspect ratio on emulsion stabilization and viscoelastic properties of the oil−water interface.11 The authors discovered that emulsion stability and the magnitude of the interfacial viscoelastic properties depend strongly on the aspect ratio of the particles.11 Strong attractive capillary interactions that are induced based on particle shape also facilitate particle assembly into networks at the oil−water interface.11 For both cylindrical and spherical particles the work required to detach a particle from the interface into a bulk phase increases from 0° to 90°, reaches a maximum at 90°, and then decreases from 90° to 180°.1 Therefore, tuning the surface chemistry of particles to intermediate wettability provides a means to prepare optimally stable particle-stabilized emulsions because of the improved propensity of the particles to partition the oil−water interface.1,9 Traditionally, hydrophilic solids can be modified by chemisorption of hydrophobic moieties such as silane coupling agents,9,12 carboxylic acids,13 and polymer grafts14−16 or physisorption of oppositely charged surfactant,17−19 polymer,20,21 and asphatlenes.22,23 Relevant to the treatment of subsea oil spills, the synergy of the hydrophobic surfactant Span 20 and hydrophilic silica particles in dispersing oil into small droplets has been demonstrated in high shear energy jet experiments.24 The adsorption of Span 20 onto the

2. EXPERIMENTAL PROCEDURES 2.1. Materials. Medium molecular weight chitosan (190−310 kDa) was obtained from Sigma-Aldrich. The reported degree of deacetylation is 75−85%. Dodecane and sodium hydroxide were obtained from Sigma-Aldrich. Acetic acid (glacial, ≥99.7%) was purchased from Fisher Scientific. All materials were used as received. Halloysite nanotubes (HNT) were purchased from NaturalNano Inc. (Rochester, NY). Deionized (DI) water produced from an Elga water purification system (Medica DV25) with a resistivity of 18.2 MΩ cm was used in all experiments. Acetic acid solutions were prepared by diluting the glacial acetic acid in DI water. 2.2. Synthesis and Characterization of Carbonized Halloysite Nanotubes (CHNT). Stock solutions of 2 wt % HNT in deionized water and 0.5 wt % medium molecular weight chitosan in acetic acid were first prepared. Appropriate amounts of medium molecular weight chitosan and HNT were then pipetted from stock solutions and added to a vial. The mass ratios of chitosan to halloysite used in particle synthesis were 0.002, 0.01, 0.02, 0.025, 0.05, 0.2, 0.5, and 1. To ensure homogeneous solubilization of chitosan and protonation of the amine groups in the solvent medium, the pH was adjusted to ∼3.7 by adding acetic acid solution. The pH was measured using a Thermo Scientific Orion 3-star benchtop pH meter. The mixture was stirred for 24 h to facilitate the adsorption of the positively charged chitosan biopolymer on the negatively charged external surface of the HNT.28 The pH was then adjusted by the addition of 1 M NaOH solution to deionize the chitosan and form a nondissolvable chitosan coating on the HNT.29 The chitosan-coated halloysite particles were recovered by centrifugation and washed twice with deionized water. The recovered samples were then pyrolyzed at 700 °C for 1 h under the flow of N2 gas to convert the chitosan adsorbed on the HNT to carbon.30 The carbonized halloysite samples are hereafter referred to as CHNT1 to CHNT8 with the numbers assigned in order of increasing level of carbonization. The morphology of the particles was characterized by field emission scanning electron microscopy (SEM, Hitachi S-4700) and transmission electron microscopy (TEM, JEOL 2010, operated at 200 kV). Energy dispersive X-ray spectroscopy (EDS) was carried out on the Hitachi S-4700 scanning electron microscope operating at 20 kV and a working distance of 15 mm. The sample was placed on a Cu/Zn substrate, and the EDS spectrum was acquired with the oxford INCA software. Thermogravimetric analysis (TGA) of the CHNT samples was performed in an air environment at a heating rate of 10 °C/min using a TA Instruments SDT 2960 simultaneous DTA-TGA. Similar 13701

DOI: 10.1021/acs.langmuir.5b03878 Langmuir 2015, 31, 13700−13707

Article

Langmuir to the CHNT synthesis procedure, the native HNT was heated at 700 °C for 2 h in N2 and used in the control TGA experiment. Zeta potential of halloysite−chitosan mixtures in acetic acid solution (pH ∼ 3.7) were determined by measuring the electrophoretic mobility using the phase analysis light scattering (PALS) technique (Nanobrook ZetaPALS, Brookhaven Instrument). 2 mL of the samples was transferred into a disposable polystyrene cuvette that was connected to the solvent resistant electrode, and the zeta potential was measured at 25 °C. Fourier transform infrared spectroscopy (FTIR) was performed on a Thermo Nicolet Nexus 670 FT-IR spectrometer. FTIR analysis was carried out with KBr pellets of the native HNT, chitosan, chitosan-coated HNT, and carbonized HNT. The chitosan-coated HNT was dried at 50 °C for the FTIR analysis. 2.3. Emulsion Preparation and Characterization. 1 wt % stock particle suspensions were prepared by uniformly dispersing the particles in water by magnetic stirring followed by ultrasonication (Cole-Parmer 8890) for 1 min. Appropriate amounts of the stock suspension were diluted in 20 mL glass vials containing the aqueous phase. Emulsions were prepared at dodecane to water ratio of 1:3 by vortex mixing the aqueous particle dispersion with the dodecane. Mixing was carried out for 2 min on a Thermolyne Maxi Mix II operating at 3000 rpm. The emulsion was characterized by optical microscopy and cryogenic scanning electron microscopy (cryo-SEM). In the optical microscopy imaging, a small aliquot of the emulsion was placed on a glass slide prior to imaging on a Leica DMI REZ optical microscope. The images were analyzed using Image ProPlus v. 5.0 software. CryoSEM imaging was performed using a Hitachi S-4800 field emission scanning electron microscope operated at a voltage of 3 kV and a working distance of 9 mm. The emulsion sample was first plunged into liquid nitrogen, followed by fracturing at −130 °C using a flat-edge cold knife and sublimation of the solvent at −95 °C for 5 min. The sample was sputtered with a gold−palladium composite at 10 mA for 88 s before imaging. Photographs of the emulsions were taken with a Pentax K10D digital camera. In the partitioning experiments, the particles were initially dispersed in dodecane by magnetic stirring at a 0.5 wt % particle concentration. 2 mL of the particle suspension was placed in a vial followed by the addition of 2 mL of water. The contents were mixed for 2 min (Thermolyne Maxi Mix II) at a low speed (∼1000 rpm) to prevent emulsion formation. 2.5. Interfacial Tension and Contact Angle Measurement. Interfacial tension was measured using the pendant drop method on a standard Ramé-Hart model 250 goniometer. About 15 μL of 0.01 wt % particle dispersions in water was injected into an external dodecane phase, and the drop shape was analyzed using DROPimage Advanced Software to obtain the interfacial tensions. Contact angle was measured on a Ramé-Hart contact angle goniometer using compressed disks of the particles.19,22,31 The contact angle measured on compressed disks is not fully representative of the contact angle of particles distributed on an interface. However, the method does provide an indirect estimation of the wettability and has been used to characterize particles at the oil−water interface.1,22 The particles were compressed into 10 mm disks in an evacuable pellet die (Specac) at a pressure of 30 MPa (Riken high pressure hydraulic equipment). The compressed disks were immersed in dodecane contained in a rectangular cell, and water was then injected from a 21 gauge needle onto the particle disks using the automated Ramé-Hart dispenser. The contact angle was measured at 25 °C through the aqueous phase at the dodecane, water, and particle interface.

Figure 2. Synthesis of the carbonized halloysite nanotubes (a); FTIR analysis of chitosan adsorption and carbonization mechanisms on halloysite nanotubes (b). Chitosan adsorbs on halloysite by electrostatic attraction and hydrogen bonding, yielding carbonized halloysite nanotubes on pyrolysis in an inert atmosphere.

tetrahedral sheet imparts a net negative charge on the external surface of the HNT.33 Figure 2a is a schematic of the adsorption of positively charged chitosan polymer on the negatively charged HNT yielding the carbonized HNT on pyrolysis in an inert atmosphere. We carried out FTIR to gain insights on the mechanisms of adsorption and subsequent pyrolysis of chitosan (Figure 2b). The FTIR spectrum of the chitosan shows key characteristic absorption peaks at 3421, 2928−2879, 1653, and 1078 cm−1 attributable to OH stretching, C−H stretching vibrations, amide I (−CONH−), and skeletal vibration involving the COO stretching, respectively.30,34,35 The FTIR for the native HNT shows the characteristic peaks at 1635, 1100, 1030, 912, 693, and 539 cm−1 corresponding to the O−H deformation vibration of the interlayer water, Si−O stretching vibration, in-plane Si−O−Si stretching vibration, O−H deformation of inner surface hydroxyl groups, perpendicular Si−O stretching vibration, and Al−O−Si deformation, respectively.12 The peaks at 3690 and 3620 cm−1 are assigned to the Al2−OH stretching bands of the HNT.28 The key observation in the spectrum for the chitosan/HNT system is that the amide I band of chitosan is perturbed from 1653 cm−1 to a lower frequency of 1576 cm−1, corresponding to a frequency shift (Δv) of 77 cm−1. The frequency shift is attributable to hydrogen-bonding interactions between the chitosan and the HNT.28,36 The zeta potential of chitosan/halloysite mixtures at pH of 3.7 (Table 1) transitioned from negative (−3.06 mV) to positive, reaching a relatively constant value (∼+70 mV) with increasing mass ratio of chitosan to halloysite. Thus, the

3. RESULTS AND DISCUSSION 3.1. Particle Characterization. Figure 2a shows a schematic for the synthesis of the carbonized halloysite nanotubes (CHNT). Figure 2b presents FTIR analysis of adsorption of chitosan onto halloysite and the synthesized CHNT. Chitosan is solubilized in acidic aqueous solution as a positively charged biopolymer due to the protonation of its primary amine groups below its pKa of 6.5.32 On the other hand, the isomorphic substitution of Al3+ for Si4+ in the external 13702

DOI: 10.1021/acs.langmuir.5b03878 Langmuir 2015, 31, 13700−13707

Article

Langmuir

obtained from the TGA. The level of carbonization increases with the increasing mass ratio of chitosan to HNT in the precursor solution used for synthesizing the CHNT (Figures 2a and 3). Accordingly, the particles transition from white to gray and ultimately to black in appearance with increasing levels of carbonization (Figure 3b). The mass losses observed in the thermogravimetric curves of the CHNT samples are due to the oxidation of the carbon coating on the HNT, and it increases with the level of carbonization. The SEM images in Figures 4a,b reveal the smooth external surface of the native HNT. SEM imaging provides direct

Table 1. Zeta Potential of Chitosan/Halloysite Solutions mass ratio of chitosan to halloysite

zeta potential (mV) −3.06 4.24 23.41 55.47 61.28 70.24 71.73 67.37

0 0.002 0.01 0.025 0.05 0.1 0.5 1

± ± ± ± ± ± ± ±

1.62 0.95 3.5 2.14 0.85 2.29 2.67 1.03

adsorption of chitosan onto the HNT is driven by electrostatic attraction and hydrogen-bonding interactions. Pyrolysis leads to loss of the aliphatic C−H peaks (2928−2879 cm−1) indicative of the decomposition of aliphatic structures37 and the loss of Al2−OH stretching bands in the CHNT, suggesting dehydration of the HNT on heat treatment at 700 °C. The FTIR spectra of the CHNT indicate the appearance of absorption bands at 1616 cm−1 attributable to CC stretching and 800 cm−1 region attributable to aromatic C−H out-of-plane bending vibrations.38 Figure 3 presents a photograph of the HNT with increasing levels of carbonization and representative TGA curves of the

Figure 3. Representative thermogravimetric curves (a) and photograph of HNT with increasing levels of carbonization (b).

Figure 4. SEM images of the native HNT (a, b) and carbonized halloysite nanotubes CHNT5 (c, d) and CHNT7 (e). Panel f is the TEM image showing the retention of the nanotubular structure in the carbonized halloysite (CHNT5), and panel g is the SEM EDS confirming the formation of the carbon coating on the aluminosilicate HNT.

particles. Table 2 lists the mass percent of carbon coating on halloysite nanotubes for samples reported in this work as Table 2. Thermogravimetric Analysis (TGA) of Carbonized Halloysite sample

mass % of carbon coating on halloysite nanotube

CHNT1 CHNT2 CHNT3 CHNT4 CHNT5 CHNT6 CHNT7 CHNT8

0.84 1.19 1.63 2.57 3.51 4.65 8.76 13.43

evidence of the carbon coating on the native HNT in the synthesis of the partially carbonized CHNT5 sample (Figures 4c,d) and the CHNT7 particles with a higher level of carbonization (Figure 4e). The TEM image shows that the CHNT retains an overall tubular nanostructure (Figure 4f). The carbonization procedure allows for the selective hydrophobic modification of the external surface of the HNT while preserving the native positively charged internal alumina surface and lumen volume. The available lumen volume in the clay nanotubes allows the loading and release of materials such as surfactant from the nanotubes in oil spill remediation applications.4,26 Elemental 13703

DOI: 10.1021/acs.langmuir.5b03878 Langmuir 2015, 31, 13700−13707

Article

Langmuir mapping by SEM/EDS on the CHNT7 sample confirms the formation of the carbon coating on the native aluminosilicate HNT (Figure 4). 3.2. Particle Partitioning and Oil Emulsification Characteristics. The oil−water partitioning characteristics of the HNT with varying levels of carbonization are presented in Figure 5. The native HNT partitions largely into the bottom

Figure 5. Influence of surface carbonization on the wettability and partitioning characteristics of halloysite clay nanotubes. Photographs a and b were taken immediately after gentle mixing and after 1 h, respectively.

aqueous phase (Figure 5a). The greater affinity of the native HNT for the aqueous phase compared to the oil phase indicates a relatively hydrophilic surface property. The aqueous HNT suspension is stable as shown by the turbidity of the bottom aqueous phase in the photograph taken after 1 h (Figure 5b). On the other hand, HNT with a high level of carbonization (CHNT7) is retained in the upper dodecane phase. Over time the CHNT7 particles sediment toward the oil−water interface under gravity.3 The particles are trapped at the oil−water interface and do not transfer into the bottom aqueous phase. Hydrophobic modification of the HNT by surface carbonization effectively imparts a greater affinity for the upper oil phase. The key observation is that at intermediate level of carbonization (CHNT4) the particles preferentially partition to the oil−water interface compared to either the bulk oil or water phase. The partial wetting of particles by both the oil and aqueous phases drives the location of the particles at the oil− water interface.9 Figure 6 shows optical microscopy images and photographs of emulsions stabilized by 0.1 wt % of the native HNT and the carbonized halloysite nanotubes (CHNT1 to CHNT7). The appearance of the emulsions ranges from creamy white for the native HNT to light gray for HNT with intermediate levels of carbonization and ultimately to black for emulsions prepared with HNT at high levels of carbonization. This is in line with the transition in the appearance of the particles illustrated in Figure 3b. Figure 6 presents the average droplet sizes and corresponding interfacial tension measurements for the clay nanotube laden dodecane−water interfaces. A key observation is that relative to the native HNT, the average droplet sizes decreases with increasing level of carbonization up to the CHNT4 sample and then increases at much higher levels of carbonization. The broad minimum in average droplet size is

Figure 6. (a) Optical microscopy images of dodecane-in-water emulsions stabilized by HNT with increasing level of carbonization. Particle concentrations are 0.1 wt %. Smaller droplets are obtained at intermediate levels of HNT carbonization (CHNT2, CHNT4). The insets are photographs of vials containing the emulsions. Scale bars = 100 μm. (b) Average emulsion droplet sizes and corresponding dodecane−water interfacial tensions measured using the pendant drop technique by injecting about 15 μL of 0.01 wt % particle suspensions into an external dodecane phase.

centered on the emulsion stabilized by the CHNT4 particles, having a 2.57 wt % carbon coating on the HNT surface (Table 2). Based on the oil-emulsification characteristics, the samples may be classified into four categories: the native HNT, HNT at very low level of carbonization (CHNT1), HNT at intermediate levels of carbonization (CHNT2−CHNT5), and HNT at high level of carbonization (CHNT6−CHNT8). Figure 6b indicates that the dodecane−water interfacial tension does not change significantly with the varying level of HNT carbonization. For the same amount of energy input and comparable interfacial tension values for all the particle types, the smaller average droplet sizes obtained at intermediate levels of halloysite carbonization can be attributable to an increased partitioning of the particles to the oil−water interface.1,39,40 The enhanced interfacial activity of the particles at intermediate 13704

DOI: 10.1021/acs.langmuir.5b03878 Langmuir 2015, 31, 13700−13707

Article

Langmuir

The cryo-SEM imaging (Figure 8) shows that networks of the CHNT are adsorbed at the oil−water interface. The CHNT

level of carbonization necessitates the creation of more oil− water interfacial area in form of small droplets to accommodate the particles at the interface.1 Once the particles are attached to the oil−water interface, they provide steric hindrance to the coalescence of the small droplets into larger ones. Figure 7a presents the variation of the dodecane−water− particle contact angles with the level of HNT carbonization.

Figure 7. (a) Variation of the three-phase contact angle with level of HNT carbonization. Red symbols are contact angle measurements while the blue curve with blue symbols corresponds to the mass percentage of carbon coating on halloysite. (b) Photographs showing contact angle of water drops on particle surfaces in an external dodecane phase.

Figure 8. Cryo-SEM images of dodecane-in-water emulsion stabilized by carbonized halloysite nanotubes (CHNT). Panels (a) to (d) are prepared with CHNT2 particles while (e) to (h) are stabilized by CHNT6 particles with a higher level of carbonization. The emulsion is stabilized by networks of the nanotubes adsorbed at the oil water interface. Particle concentration is 0.1 wt% in water.

The three-phase contact angle for the CHNT4 sample is 89.8°. This contact angle value is the closest to 90° of all particle types and coincides with the minimum in average droplet sizes centered on the emulsions stabilized by the CHNT4 (Figure 6). However, we note that the compressed disk method is an indirect way to characterize the change in hydrophobicity of the nanotubes. The representative photographs in Figure 7b illustrate the variation of the three-phase contact angle with increasing level of HNT carbonization. Regression analysis shows that the contact angle (θ) varies logarithmically with the level of HNT carbonization (mc) according to the equation θ = 37.27 ln(mc) + 54.52

stabilizes the oil droplets by providing a steric barrier to droplet coalescence due to the high energy of attachment of the particles at the oil−water interface.4 Coalescence of solidstabilized emulsion droplets can occur when the particles are detached from the interface into a bulk phase or the particles are displaced laterally along the interface. The lateral displacement of the particles at the oil−water interface could result in the contact of exposed oil−water interfacial areas on adjacent droplets leading to coalescence.1 In particle-stabilized emulsions, the strength of the particle network adsorbed at the oil− water interface prevents lateral displacement of particles away from droplet contact areas, providing stability against coalescence. 1 The energy required to detach network aggregates of particles at the oil−water interface is also much higher than for an individual particle as it scales with the overall size of the adsorbed particle aggregates (eqs 1−3).2,4 These mechanisms are fundamental to emulsion stabilization by solids such as the CHNT.

(4) 2

with the coefficient of determination (R ) = 0.97. At HNT surface carbonization (mc) levels of 0.84, 2.57, and 13.43 wt %; the contact angles which the particle makes at the oil−water interface are calculated to be 48.0°, 89.7°, and 151.3°, respectively. Thus, carbonization of the external surface of the native hydrophilic halloysite particles effectively tunes the wettability of the particles for emulsion stabilization. 13705

DOI: 10.1021/acs.langmuir.5b03878 Langmuir 2015, 31, 13700−13707

Article

Langmuir

(6) Athas, J. C.; Jun, K.; McCafferty, C.; Owoseni, O.; John, V. T.; Raghavan, S. R. An Effective Dispersant for Oil Spills Based on FoodGrade Amphiphiles. Langmuir 2014, 30 (31), 9285−9294. (7) Venkataraman, P.; Tang, J. J.; Frenkel, E.; McPherson, G. L.; He, J. B.; Raghavan, S. R.; Kolesnichenko, V.; Bose, A.; John, V. T. Attachment of a Hydrophobically Modified Biopolymer at the OilWater Interface in the Treatment of Oil Spills. ACS Appl. Mater. Interfaces 2013, 5 (9), 3572−3580. (8) Owoseni, O.; Nyankson, E.; Zhang, Y.; Adams, D. J.; He, J.; Spinu, L.; McPherson, G. L.; Bose, A.; Gupta, R. B.; John, V. T. Interfacial adsorption and surfactant release characteristics of magnetically functionalized halloysite nanotubes for responsive emulsions. J. Colloid Interface Sci. 2016, 463, 288−298. (9) Binks, B. P.; Lumsdon, S. O. Influence of particle wettability on the type and stability of surfactant-free emulsions. Langmuir 2000, 16 (23), 8622−8631. (10) Creighton, M. A.; Ohata, Y.; Miyawaki, J.; Bose, A.; Hurt, R. H. Two-Dimensional Materials as Emulsion Stabilizers: Interfacial Thermodynamics and Molecular Barrier Properties. Langmuir 2014, 30 (13), 3687−3696. (11) Madivala, B.; Vandebril, S.; Fransaer, J.; Vermant, J. Exploiting particle shape in solid stabilized emulsions. Soft Matter 2009, 5 (8), 1717−1727. (12) Yuan, P.; Southon, P. D.; Liu, Z. W.; Green, M. E. R.; Hook, J. M.; Antill, S. J.; Kepert, C. J. Functionalization of halloysite clay nanotubes by grafting with gamma-aminopropyltriethoxysilane. J. Phys. Chem. C 2008, 112 (40), 15742−15751. (13) Zhou, J.; Wang, L. J.; Qiao, X. Y.; Binks, B. P.; Sun, K. Pickering emulsions stabilized by surface-modified Fe3O4 nanoparticles. J. Colloid Interface Sci. 2012, 367, 213−224. (14) Foster, L. M.; Worthen, A. J.; Foster, E. L.; Dong, J. N.; Roach, C. M.; Metaxas, A. E.; Hardy, C. D.; Larsen, E. S.; Bollinger, J. A.; Truskett, T. M.; Bielawski, C. W.; Johnston, K. P. High Interfacial Activity of Polymers “Grafted through” Functionalized Iron Oxide Nanoparticle Clusters. Langmuir 2014, 30 (34), 10188−10196. (15) Hou, Y. F.; Jiang, J. Q.; Li, K.; Zhang, Y. W.; Liu, J. D. Grafting Amphiphilic Brushes onto Halloysite Nanotubes via a Living RAFT Polymerization and Their Pickering Emulsification Behavior. J. Phys. Chem. B 2014, 118 (7), 1962−1967. (16) Lu, J.; Zhou, W.; Chen, J.; Jin, Y. L.; Walters, K. B.; Ding, S. J. Pickering emulsions stabilized by palygorskite particles grafted with pH-responsive polymer brushes. RSC Adv. 2015, 5 (13), 9416−9424. (17) Binks, B. P.; Rodrigues, J. A.; Frith, W. J. Synergistic interaction in emulsions stabilized by a mixture of silica nanoparticles and cationic surfactant. Langmuir 2007, 23 (7), 3626−3636. (18) Zhu, Y.; Jiang, J. Z.; Liu, K. H.; Cui, Z. G.; Binks, B. P. Switchable Pickering Emulsions Stabilized by Silica Nanoparticles Hydrophobized in Situ with a Conventional Cationic Surfactant. Langmuir 2015, 31 (11), 3301−3307. (19) Cavallaro, G.; Lazzara, G.; Milioto, S.; Parisi, F. Hydrophobically Modified Halloysite Nanotubes as Reverse Micelles for Water-in-Oil Emulsion. Langmuir 2015, 31 (27), 7472−7478. (20) Skelhon, T. S.; Grossiord, N.; Morgan, A. R.; Bon, S. A. F. Quiescent water-in-oil Pickering emulsions as a route toward healthier fruit juice infused chocolate confectionary. J. Mater. Chem. 2012, 22 (36), 19289−19295. (21) Williams, M.; Armes, S. P.; York, D. W. Clay-Based Colloidosomes. Langmuir 2012, 28 (2), 1142−1148. (22) Yan, N. X.; Masliyah, J. H. Effect of pH on adsorption and desorption of clay particles at oil-water interface. J. Colloid Interface Sci. 1996, 181 (1), 20−27. (23) Jada, A.; Debih, H. Hydrophobation of Clay Particles by Asphaltenes Adsorption. Compos. Interfaces 2009, 16 (2−3), 219−235. (24) Yu, G. Z.; Dong, J. N.; Foster, L. M.; Metaxas, A. E.; Truskett, T. M.; Johnston, K. P. Breakup of Oil Jets into Droplets in Seawater with Environmentally Benign Nanoparticle and Surfactant Dispersants. Ind. Eng. Chem. Res. 2015, 54 (16), 4243−4251. (25) Cavallaro, G.; Lazzara, G.; Milioto, S.; Parisi, F.; Sanzillo, V. Modified Halloysite Nanotubes: Nanoarchitectures for Enhancing the

4. CONCLUSIONS Naturally occurring halloysite nanotubes have been hydrophobized by the selective carbonization of the external surface of the HNT. The level of carbonization of the HNT determines the relative wettability of the particles by the oil or water phases. Smaller droplet sizes are obtained at intermediate levels of carbonization compared to the hydrophilic native HNT due to an increased propensity of the particles to partition to the oil−water interface. The lowest average droplet sizes are obtained for emulsions stabilized by the CHNT4 particles prepared at a chitosan to halloysite mass ratio of 0.025, yielding a carbon coating of about 2.57 wt %. Droplet size analysis indicates that at high levels of carbonization the HNT are less effective in oil emulsification into small droplets. The driving force for oil emulsification into smaller droplets at intermediate levels of carbonization is the improved preference of the particles to reside at the oil−water interface. The experimental observations with increasing levels of HNT carbonization are in agreement with free energy analysis on the attachment of cylindrical particles at the oil−water interface. The energy required to detach a cylindrical particle from the oil−water interface is much higher at intermediate particle wettability compared to the more hydrophilic or hydrophobic regimes. The concept of tuning the wettability of solids by surface carbonization can be extended to other inorganic materials such as titania or iron oxides as well as other carbon sources including sugars, polymers, and organic molecules. The lumen volume in the clay nanotubes can be exploited for the smart delivery of materials such as surfactant,4,26 fluorescent markers, or nutrients in oil spill remediation applications.41



AUTHOR INFORMATION

Corresponding Author

*E-mail [email protected]; Ph (504) 865-5883; Fax (504) 8656744 (V.T.J.). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Funding from the Gulf of Mexico Research Initiative with additional support from the Louisiana Board of Regents and the National Science Foundation (Grant 1236089) is gratefully acknowledged.



REFERENCES

(1) Binks, B. P., Horozov, T. S., Eds.; Colloidal Particles at Liquid Interfaces; Cambridge University Press: Cambridge, UK, 2006. (2) Saha, A.; Nikova, A.; Venkataraman, P.; John, V. T.; Bose, A. Oil Emulsification Using Surface-Tunable Carbon Black Particles. ACS Appl. Mater. Interfaces 2013, 5 (8), 3094−3100. (3) Dong, J. N.; Worthen, A. J.; Foster, L. M.; Chen, Y. S.; Cornell, K. A.; Bryant, S. L.; Truskett, T. M.; Bielawski, C. W.; Johnston, K. P. Modified Montmorillonite Clay Microparticles for Stable Oil-inSeawater Emulsions. ACS Appl. Mater. Interfaces 2014, 6 (14), 11502− 11513. (4) Owoseni, O.; Nyankson, E.; Zhang, Y. H.; Adams, S. J.; He, J. B.; McPherson, G. L.; Bose, A.; Gupta, R. B.; John, V. T. Release of Surfactant Cargo from Interfacially-Active Halloysite Clay Nanotubes for Oil Spill Remediation. Langmuir 2014, 30 (45), 13533−13541. (5) Worthen, A. J.; Foster, L. M.; Dong, J.; Bollinger, J. A.; Peterman, A. H.; Pastora, L. E.; Bryant, S. L.; Truskett, T. M.; Bielawski, C. W.; Johnston, K. P. Synergistic formation and stabilization of oil-in-water emulsions by a weakly interacting mixture of zwitterionic surfactant and silica nanoparticles. Langmuir 2014, 30 (4), 984−94. 13706

DOI: 10.1021/acs.langmuir.5b03878 Langmuir 2015, 31, 13700−13707

Article

Langmuir Capture of Oils from Vapor and Liquid Phases. ACS Appl. Mater. Interfaces 2014, 6 (1), 606−612. (26) Nyankson, E.; Owoseni, O.; John, V. T.; Gupta, R. B. SurfactantLoaded Halloysite Clay Nanotube Dispersants for Crude Oil Spill Remediation. Ind. Eng. Chem. Res. 2015, 54 (38), 9328−9341. (27) Levis, S. R.; Deasy, P. B. Use of coated microtubular halloysite for the sustained release of diltiazem hydrochloride and propranolol hydrochloride. Int. J. Pharm. 2003, 253 (1−2), 145−157. (28) Liu, M. X.; Zhang, Y.; Wu, C. C.; Xiong, S.; Zhou, C. R. Chitosan/halloysite nanotubes bionanocomposites: Structure, mechanical properties and biocompatibility. Int. J. Biol. Macromol. 2012, 51 (4), 566−575. (29) Liu, Y. Y.; Tang, J.; Chen, X. Q.; Xin, J. H. Decoration of carbon nanotubes with chitosan. Carbon 2005, 43 (15), 3178−3180. (30) Bengisu, M.; Yilmaz, E. Oxidation and pyrolysis of chitosan as a route for carbon fiber derivation. Carbohydr. Polym. 2002, 50 (2), 165−175. (31) Abdullayev, E.; Price, R.; Shchukin, D.; Lvov, Y. Halloysite Tubes as Nanocontainers for Anticorrosion Coating with Benzotriazole. ACS Appl. Mater. Interfaces 2009, 1 (7), 1437−1443. (32) Chen, Y. J.; Javvaji, V.; MacIntire, I. C.; Raghavan, S. R. Gelation of Vesicles and Nanoparticles Using Water-Soluble Hydrophobically Modified Chitosan. Langmuir 2013, 29 (49), 15302−15308. (33) Joussein, E.; Petit, S.; Churchman, J.; Theng, B.; Righi, D.; Delvaux, B. Halloysite clay minerals - A review. Clay Miner. 2005, 40 (4), 383−426. (34) Chen, L.; Tang, C. Y.; Ning, N. Y.; Wang, C. Y.; Fu, Q.; Zhang, Q. Preparation and Properties of Chitosan/Lignin Composite Films. Chin. J. Polym. Sci. 2009, 27 (5), 739−746. (35) Ding, Y.; Gu, G.; Xia, X. H.; Huo, Q. Cysteine-grafted chitosanmediated gold nanoparticle assembly: from nanochains to microcubes. J. Mater. Chem. 2009, 19 (6), 795−799. (36) Hertl, W. Infrared frequency shifts due to hydrogen bonding of surface amino groups on silica. J. Phys. Chem. 1973, 77, 1473. (37) Kaczmarek, H.; Zawadzki, J. Chitosan pyrolysis and adsorption properties of chitosan and its carbonizate. Carbohydr. Res. 2010, 345 (7), 941−947. (38) Sevilla, M.; Fuertes, A. B. The production of carbon materials by hydrothermal carbonization of cellulose. Carbon 2009, 47 (9), 2281− 2289. (39) Yan, N. X.; Masliyah, J. H. Characterization and Demulsification of Solids-Stabilized Oil-in-Water Emulsions 0.1. Partitioning of Clay Particles and Preparation of Emulsions. Colloids Surf., A 1995, 96 (3), 229−242. (40) Schulman, J. H.; Leja, J. Control of contact angles at the oil− water−solid interfaces. Trans. Faraday Soc. 1954, 50, 598−605. (41) Fingas, M. Oil Spill Science and Technology; Gulf Professional Publishing: Boston, MA, 2011.

13707

DOI: 10.1021/acs.langmuir.5b03878 Langmuir 2015, 31, 13700−13707